Vous êtes sur la page 1sur 20

Review of models of mode

instability in fiber amplifiers


Arlee V. Smith and Jesse J. Smith
AS-Photonics, LLC, 8500 Menaul Blvd. NE, Suite B335, Albuquerque, NM
87112 USA
arlee.smith@as-photonics.com

Abstract: We compare several published models of mode instabil-


ity for fiber amplifiers.
2013 Optical Society of America
OCIS codes: (060.2320) Fiber optics amplifiers and oscillators; (060.4370) Non-
linear optics, fibers; (140.6810) Thermal effects; (190.2640) Stimulated scattering,
modulation, etc

References and links


1. T. Eidam, C. Wirth, C. Jauregui, F. Stutzki, F. Jansen, H.-J. Otto, O. Schmidt, T. Schreiber,
J. Limpert, and A. T unnermann, Experimental observations of the threshold-like onset
of mode instabilities in high power fiber amplifiers, Opt. Exp. 19, 23218-13224 (2011),
http://www.opticsinfobase.org/oe/abstract.cfm?URI=oe-19-14-13218.
2. M. Karow, H. T unnermann, J. Neumann, D. Kracht, and P. Wessels, Beam
quality degradation of a single-frequency Yb-doped photonic crystal fiber am-
plifier with low mode instability threshold, Opt. Lett. 37, 42424244 (2012),
http://www.opticsinfobase.org/oe/abstract.cfm?URI=ol-37-20-4242.
3. H.-J. Otto, F. Stutzki, F. Jansen, T. Eidam, C. Jauregui, J. Limpert, and A. T unnermann,
Temporal dynamics of mode instabilities in high-power fiber lasers and amplifiers, Opt. Express
20, 1571015722 (2012), http://www.opticsinfobase.org/oe/abstract.cfm?URI=oe-20-14-15710.
4. B. Ward, C. Robin, and I. Dajani, Origin of thermal modal instabilities
in large mode area fiber amplifiers, Opt. Express 20, 1140711422 (2012),
http://www.opticsinfobase.org/oe/abstract.cfm?URI=oe-20-10-11407.
5. F. Stutzki, H.-J. Otto, F. Jansen, C. Gaida, C. Jauregui, J. Limpert, and A. T
unnermann, High-
speed modal decomposition of mode instabilities in high-power fiber lasers, Opt. Lett. 36, 4572-
4574 (2011), http://www.opticsinfobase.org/ol/abstract.cfm?URI=ol-36-23-4572.
6. A. V. Smith and J. J. Smith, Mode instability in high power fiber amplifiers, Opt. Express 19,
1018010192 (2011), http://www.opticsinfobase.org/oe/abstract.cfm?URI=oe-19-11-10180.
7. A. V. Smith and J. J. Smith, Influence of pump and seed modulation on the
mode instability thresholds of fiber amplifiers, Opt. Express 20, 2454524558 (2012),
http://www.opticsinfobase.org/oe/abstract.cfm?URI=oe-20-22-24545.
8. A. V. Smith and J. J. Smith, Steady-periodic method for modeling
mode instability in fiber amplifiers, Opt. Express 21, 2606-2623 (2013)
http://www.opticsinfobase.org/oe/abstract.cfm?URI=oe-21-3-2606.
9. A. V. Smith and J. J. Smith, Maximizing the mode instability threshold of a fiber amplifier,
arXiv:1301.3489 [physics.optics] (2013), http://arxiv.org/abs/1301.3489.
10. A. V. Smith and J. J. Smith, Frequency dependence of mode coupling gain in Yb doped fiber
amplifiers due to stimulated thermal Rayleigh scattering,arXiv:1301.4277 [physics.optics] (2013),
http://arxiv.org/abs/1301.4277.
11. A. V. Smith and J. J. Smith, Modeled fiber amplifier performance near the mode instability
threshold, arXiv:1301.4278 [physics.optics] (2013), http://arxiv.org/abs/1301.4278.
12. A. V. Smith and J. J. Smith, Increasing mode instability thresholds of
fiber amplifiers by gain saturation, Opt. Express 21, 15168-15182 (2013),
http://www.opticsinfobase.org/oe/abstract.cfm?URI=oe-21-13-15168.
13. A. V. Smith and J. J. Smith, Spontaneous Rayleigh seeding of stimulated Rayleigh
scattering in high power fiber amplifiers, arXiv:1301.2763 [physics.optics] (2013),
http://arxiv.org/abs/1301.2763.
14. K. R. Hansen, T. T. Alkeskjold, J. Broeng, and J. Laegsgaard, Thermally induced
mode coupling in rare-earth doped fiber amplifiers, Opt. Lett. 37, 23822384 (2012),
http://www.opticsinfobase.org/oe/abstract.cfm?URI=ol-37-12-2382.
15. K. R. Hansen, T. T. Alkeskjold, J. Broeng, and J. Lgsgaard, Theoretical analysis
of mode instability in high-power fiber amplifiers, Opt. Express 21, 19441971 (2013),
http://www.opticsexpress.org/abstract.cfm?URI=oe-21-2-1944.
16. L. Dong, Stimulated thermal rayleigh scattering in optical fibers, Opt. Express 21, 26422656
(2013), http://www.opticsexpress.org/abstract.cfm?URI=oe-21-3-2642.
17. C. Jauregui, T. Eidam, H.-J. Otto, F. Stutzki, F. Jansen, J. Limpert, and A. T unnermann,
Physical origin of mode instabilities in high-power fiber laser systems, Opt. Express 20, 12912-
12925 (2012), http://www.opticsinfobase.org/oe/abstract.cfm?URI=oe-20-12-12912.
18. I-Ning Hu, C. Zhu, C. Zhang, A. Thomas, and A. Galvanauskas, Analytical time-dependent
theory of thermally-induced modal instabilities in high power fiber amplifiers, Photonics West
Conf. Fiber Lasers X, paper 8601-109 (2013)
19. M. D. Feit and J. A. Fleck, Computation of mode properties in optical fiber
waveguides by a propagating beam method, Applied Optics 19, 1154-1164 (1980),
http://ao.osa.org/abstract.cfm?URI=ao-19-7-1154.
20. M. D. Feit and J. A. Fleck, Computation of mode eigenfunctions in graded-index op-
tical fibers by the propagating beam method, Applied Optics 19, 2240-2246 (1980),
http://ao.osa.org/abstract.cfm?URI=ao-19-13-2240.
21. K. D. Cole, Steady-periodic Greens functions and thermal-measurement applications in rectan-
gular coordinates, J. of Heat Trans. 128, 706-716 (2006); DOI: 10.1115/1.2194040
22. K. D. Cole and P. E. Crittenden, Steady-periodic Heating of a cylinder, J. of Heat Trans. 131,
091301-1091301-7 (2009); DOI: 10.1115/1.3139107
23. S. Naderi, I. Dajani, T. Madden, B. Ward, C. Robin, and J. Grosek, Numerical studies of modal
instabilities in high-power fiber amplifiers, Photonics West Conf. Fiber Lasers X, paper 8601-125
(2013).
24. C. Jauregui, J. Limpert, and A. T unnermann, Origin of thermal modal instabil-
ities in large mode area fiber amplifiers, Spotlight on Optics, June 11, 2012,
http://www.opticsinfobase.org/spotlight/summary.cfm?URI=oe-20-10-11407.
25. S. Naderi, I. Dajani, T. Madden, and C. Robin, Investigations of modal instabilities in
fiber amplifiers through detailed numerical simulations, Opt. Exp. 21, 16111-16129 (2013),
http://www.opticsexpress.org/abstract.cfm?URI=oe-21-13-16111.
26. B.G. Ward, Modeling of transient modal instability in fiber amplifiers, Opt. Exp. 21, 12053-
12067 (2013), http://www.opticsexpress.org/abstract.cfm?URI=oe-21-10-12053.
27. M.M. Johansen, K.R. Hansen, M. Laurila, T.T. Alkeskjold, and J. Laegsgaard, Estimating
modal instability threshold for photonic crystal rod fiber amplifiers, Opt. Exp. 21, 15409-15417,
http://www.opticsexpress.org/abstract.cfm?URI=oe-21-13-15409.

1. Introduction
Mode instability refers to the threshold-like degradation of the output beam profile
from large mode area fiber amplifiers. The instability has been observed by several
researchers [15] in Yb-doped amplifiers pumped at 976 nm and operating in the range
1030-1080 nm. Reported instability thresholds fall in the range of 100 W to 2500 W.
The instability has been attributed to stimulated thermal Rayleigh scattering (STRS)
by some [612, 1416, 18] and to perhaps alternative thermal mode coupling processes
by others [4, 17, 23, 25, 26].
In all models the mode coupling process responsible for the instability is thermal.
Regions of the core are heated by the quantum defect fraction of the absorbed pump
light. The resulting temperature profile creates a refractive index grating via the thermo
optic effect, and this index grating scatters light from mode LP01 into higher order
modes, primarily into LP11 . In all the models that are compared here, the critical
model components are calculations of the quantum defect heating associated with signal
amplification, calculations of the resulting temperature and refractive index profiles, and
calculations of the rate of power transfer from LP01 to LP11 . It is generally recognized
that such mode coupling requires a phase shift between the irradiance grating produced
by modal interference and the index grating it creates. However, the cause of that
phase shift is assumed to be a frequency offset between modes in the STRS models
[612, 1416, 18], and perhaps to other effects in the remaining models [4, 17, 23, 25, 26].
In the following sections we describe the philosophies and the approximations of each
model. When possible we compare predictions of the different models with one another
and with measured performance of laboratory amplifiers.

2. Model approximations and methods


All physical models necessarily involve numerous approximations of the actual physics.
The key to a useful model is selecting the appropriate approximations for the applica-
tion of interest. Two different philosophies have been adopted based on two different
assumptions about the cause of the instability. One model type starts with a transient
model and runs to convergence. The other model type starts with the steady-periodic
assumption and integrates over one period of the beat between two optical fields that
are offset in frequency. In this type model a small perturbation is added to the steady
state solution of amplification to determine whether it grows exponentially and, if so,
whether threshold is reached. The physics of heating and mode coupling is nearly the
same in the two model types, so we expect they would converge if periodic fields were
used as inputs to the transient model.
Other choices that differ among models include the method of beam propagation and
the method of solving the thermal diffusion equation. A beam propagation model (BPM)
includes all transverse modes automatically. A coupled mode model requires that the
user specify which modes to include as well as defining the modes. A large number of
included modes would require integrating a large number of mode coupling equations.
The steady-periodic assumption allows the use of steady-periodic Greens functions to
solve the temperature equation. Alternatively, Dong [16] deals with the heat equation by
describing the temperature in terms of a set of temperature modes derived in cylindrical
coordinates. The alternating direction explicit/implicit (ADI) method can also be used
to solve the temperature equation.
Below we list some of the approximations that have been used in mode instability
models. The choice of approximations can strongly influence the computational speed,
with some choices being advantageous in exploiting parallel computing methods.

2.1. Number of modes included


A coupled mode treatment requires explicit inclusion of all modes of interest. A coupling
equation is added to the model for each coupled mode pair. If only two modes are
of interest this can be an advantage, but if many modes are important, it can be a
disadvantage. Usually only two modes are of primary interest in computing thresholds,
but for above threshold operation many high order modes might be important.
A beam propagation model (BPM) automatically includes all modes because a single
signal field is propagated without regard to modal content.
2.2. Thermal boundary condition
There are several choices to make for the time dependent thermal model. The first is
whether to use a two dimensional (x, y,t) or three dimensional (x, y, z,t) solver. A two
dimensional solver is applied to each z position, independent of adjacent z positions.
The two dimensional solver is probably faster than the three dimensional solver, but
it does not allow longitudinal flow of heat. The second choice is what to use as the
thermal boundary condition. If the average temperature of the core, rather than the
much smaller temperature variation near the core, is expected to be important, the
model boundary should coincide with the outer diameter of the fiber. If a polymer
coating is used, it should be included. A further complication may arise if the fiber
under simulation is a photonic crystal fiber (that is, it has a complicated structure of
thermal conductivity and density). If slow transient effects are of interest the model
boundary should match the physical boundary.
However, if only transverse temperature variations about the average and with a
length scale comparable to the core diameter are expected to be important, a boundary
removed from the fiber center by a few times the core diameter can be used. Based
on thermal diffusion rates, frequencies of a few hundred hertz to a few kilohertz imply
length scales of a few tens of microns. If the steady-periodic assumption is imposed, the
closer boundary can be used. Fast temperature oscillations within the core region are
not affected by a nearby boundary because the heating is expected to have a frequency
corresponding to thermal diffusion across the core.
A third choice for the model boundary is its shape. It should be round to best match
the physical boundary if the full fiber is to be included. However, a square boundary at
approximately the same distance may be acceptable and may allow much faster solution.
Similarly, if only the core region is to be included, a square boundary may allow fast
solution with acceptable error.
A fourth choice is whether the boundary is held at a fixed temperature or whether
some other thermal boundary condition is imposed. If the fiber is cooled asymmetrically,
it may be more realistic to fix the temperature only on the cooled portion and allow
the other portions to float, meaning the temperature slope is fixed and small so there
is low heat flow there.

2.3. Spatial grid


Models can be based on a rectangular or a cylindrical coordinate system. Both coupled
mode and BPM modeling is possible using cylindrical coordinates. Cylindrical coor-
dinates fit better to a cylindrical thermal boundary. However, if only r dependence is
included, the optical field is limited to LP0n modes. The LP11 mode that is usually
observed in the laboratory is not accommodated. Of course, if the dependence is in-
cluded, all modes can be treated. Rectangular coordinates also permit all modes, and
they allow BPM using fast Fourier transform (FFT) methods [19, 20]. If necessary a
cylindrical thermal boundary can be step wise approximated on a rectangular grid in
some temperature solvers.

2.4. Heat deposition


Some models assume the heat deposition profile matches the signal irradiance profile
[1416]. They assume that when the signal irradiance grows by an amount I(x, y) on
propagation by z, the heat deposition is equal to I(x, y) multiplied by the quantum
defect factor. This means the profile of deposited heat matches the signal irradiance
profile within a multiplication factor. Other models [4, 612, 23, 25, 26] calculate the
population inversion [nu (x, y) nl (x, y)] at each z location and use that to compute the
absorbed pump power over one z step, assuming the pump power is uniformly distributed
over the pump cladding. The profile of the deposited heat computed from the pump
absorption is proportional to [ pa nl (x, y) pe nu (x, y)] which can strongly differ from the
signal irradiance profile if the population inversion is depleted by transverse spatial hole
burning, as was shown in [10,12]. It is an important point that efficient fiber amplifiers,
operating above a few watts, generally have strong transverse hole burning at some or
all positions along the fiber.
If photodarkening or linear absorption is included in the model, an absorption profile
(x, y, z) can be added, with the assumption that all absorbed light is converted to
heat. This can increase the mode coupling gain. Photodarkening is usually claimed to
be strongest in regions of high population inversion and to change slowly in time, so
a photodarkening contribution to (x, y, z) might be computed from the time-averaged
local population inversions.

2.5. Method of propagating beam


If a rectangular grid is used, and if the amplifier is an index guiding fiber rather than
a photonic fiber, FFT methods can be used to propagate the signal field [19, 20]. Al-
ternatively, if cylindrical symmetry is assumed, fast Hankel transforms can be used to
propagate the field. If a photonic crystal fiber (PCF) is used, it can be approximated as
a step index fiber and FFT or Hankel methods can be used, or the actual index profile
can be accommodated in a finite difference beam propagation model. Propagation using
finite element methods is also possible. The finite element and finite difference methods
have the advantage that a conformal grid can be used to handle the air holes without
a high density grid. Full vectorial FFT methods that can handle the large index steps
of a PCF are also available, but the square gridding means a dense grid is necessary to
properly account for the small air holes. Recently Ward [26] reported a beam propa-
gation method that is a hybrid of finite element method, finite differences, and Fourier
angular expansion.
However, a BPM model is probably slower than a coupled mode model because its z
step must be a small fraction of the modal beat length, usually a few microns, whereas
several times larger z steps are usually possible in coupled mode models.
Coupled mode models do not propagate fields directly. They require integration in z
of a set of equations describing the gain of each fixed profile mode due to four processes,
laser gain, thermally induced self phase modulation, thermally induced self phase mod-
ulation, and interference induced mode coupling. One equation is necessary for each
mode involved, and it includes coupling terms between each of the modes included in-
cluding itself. Coupled mode models force the light to remain in only those modes that
are specifically included. One consequence is that if only LP01 and LP11 are included,
LP01 isnt allowed to thermally lense into a profile that includes LP02 or other radially
symmetric modes. Additionally, for operation above threshold where a large number of
modes may be important, will not be realistically modeled. The advantage of a coupled
mode treatment is that the z-step can be larger than in a BPM model. This is a limited
advantage because the thermal equation must be solved just as often for either model,
and the thermal solver usually occupies substantially more time than the beam prop-
agation model, even in the BPM models. Another advantage of coupled modes is that
the mode profiles can be those of a PCF or other complex index structure, whereas the
modes for a split-step FFT BPM model are usually those of an index guiding fiber with
low index contrast.

2.6. Temperature solver


The choice of solution method for the time-dependent temperature equation may be the
most important choice in a mode instability model because most of the computation
time is usually spent on this task. The choice depends largely on whether the model uses
the steady-periodic assumption or not. If so, steady-periodic Greens function solutions
are available for a rectangular grid and rectangular thermal boundary conditions [21],
and for cylindrical grid and cylindrical boundary conditions [22]. Additionally, Dong [16]
developed a steady-periodic temperature solver based on temperature modes.
Other methods are available for transient models, including relaxation methods which
involve a relatively simple finite-difference explicit integration in time. Relaxation meth-
ods are general and can be used in arbitrary boundary conditions, but are very slow
because they are limited in the size of time step. Paradoxically, larger time steps
are allowed only in cases where the transverse spatial step is large (i.e. fine struc-
ture is unimportant). Implicit methods are faster. They are unconditionally stable, so
larger time steps are possible, but accuracy suffers if the time step is too large. The
alternating-direction implicit method (ADI) combines explicit and implicit integration
to improve accuracy and maintain unconditional stability. Finite-difference and finite-
element methods can also be used. They are computationally intensive, but general
enough to handle arbitrary boundary conditions and conformal meshes.

2.7. Feedback effects


A physical system with gain, time delay, and negative feedback can form an oscillator
if the gain is sufficient. The oscillation frequency is determined by the time delay.
Fiber amplifiers probably do not form such oscillators, but the presence of time delayed
feedback combined with amplifier gain could have an influence on the amplifier stability.
In the case of a fiber amplifier the feedback would be caused by changes in the signal
or pump power near the output end of the fiber leading to altered conditions earlier
in the fiber. This can happen only in counter-pumped amplifiers where the pump light
can carry feedback information toward the signal input end.
For example, if the pump absorption is smaller when LP11 is populated near the
output end, the pump power early in the fiber is increased when mode switching into
LP11 begins. This would tend to move the grating upstream, decreasing the power
transfer into LP11 , a negative feedback response. It would be strongest in fibers with
confined doping where LP01 overlaps the dopant profile better than LP11 . However,
there is also a response with the opposite sign; higher pump power early in the fiber
tends to increase the degree of power switching if the grating motion is not altered,
as would be the case near the signal input end. The net effect is unclear. It may be
that feedback plays a role in the large excursions in modal power content sometimes
observed in counter-pumped amplifiers operated above threshold [1, 3]. However, mode
instability has been documented in both co- and counter-pumped amplifiers at similar
powers, and confined doping is known to raise the threshold, so feedback must play a
secondary role in modal instability.

2.8. Slinky effects


This effect is named for the toy coil spring that flip-flops its way down stairways. Imagine
the coils representing the relative phases of the two interacting modes. In a transient
model of mode coupling, when the power balance of the two modes changes, the profile of
the temperature in the core changes in a way that either pulls the propagation constants
of the two modes closer together, in which case the slinky expands, or pushes them
farther apart, in which case the slinky contracts. The thermal grating lags the moving
light grating and this produces the phase shift that allows power transfer between the
modes. The direction of power transfer depends on the direction of movement of the local
grating or slinky coils. If motion is slow enough that the phase lag between irradiance
and temperature gratings is less than 90 , downstream motion leads to transfer power
into LP11 ; upstream motion leads to transfer power into LP01 . Slinky motions can be
complicated with upstream motion in one zone and downstream motion in another.

2.9. Thermal effects


All current models of mode instability include quantum defect heating and use it to
compute the temperature profile and from that the refractive index changes due to the
thermo optic effect. They all ignore other temperature dependent effects such as the
temperature dependence of the Yb absorption and emission cross sections, or thermally-
induced strain leading to changes in the refractive index due to the photo elastic effect.
Temperature variation of thermal conductivity and heat capacity are ignored. The ne-
glected effects are expected to be insignificant compared with the thermo optic effect.

2.10. Masking effect


When LP11 (0) is seeded with much more than the quantum noise level, it beats with
LP01 to produce a stationary irradiance and temperature grating that masks the much
weaker moving gratings produced by beating between LP01 and weak LP11 ( 6= 0) light.
The stationary gratings can be removed by subtracting the DC part of the gratings.
When this is done the weak moving gratings are sometimes clearly revealed, or un-
masked.

3. AS-Photonics Model
The earliest AS-Photonics paper [6] on mode instability describes the physics of STRS.
The numerical model we developed was presented in full detail in another paper [8].
Other papers present model results from studies of various physical effects and operating
conditions [7, 913].

Assumptions: Refractive index guiding - usually a step index, steady-periodic heating,


heating computed from the absorbed pump using the (x, y) dependent population
inversion, no longitudinal heat flow, square thermal boundary, square grid. The
steady-periodic assumption implies the frequency spectrum consists of regularly
spaced frequencies with a spacing equal to the inverse of the period.
Methods: Greens function steady-periodic temperature solver, FFT beam propaga-
tion.
Features: Highly parallelizable in both the BPM and the thermal solver; highly nu-
merical to minimize assumptions about fiber or pump or seed; run times 15-90
minutes per meter.
Notes: In Fig. 1 (from [10]) we show the computed mode coupling gain (blue curve),
total heating (red curve), and laser gain (green curve) for a co-pumped amplifier,
computed by this model. The amplifier in this case has dcore = ddope = 80 m,
L = 1.2 m, Aclad = 60000 m2 , Ppump = 600 W, P01 = 10 W, P11 = 1 1016 W. The
shape of mode coupling gain is affected by transverse spatial hole burning. Early
in the fiber the pump is strong and the signal weak so transverse hole burning is
weak and mode coupling gain is relatively high. Late in the fiber the pump is weak
and the signal is strong so transverse hole burning is strong and mode coupling
gain is relatively weak.

300 1

250

200

Heat [norm.]
Gain [dB/m]

150 0.5

100

50

0 0
0 0.2 0.4 0.6 0.8 1 1.2
Z [m]

Fig. 1. Heat (red), mode coupling gain (blue), and laser gain (green) for a co-
pumped amplifier operating near the mode instability threshold.

We further explore the effect of transverse hole burning on the frequency of max-
imum gain in [10]. The frequency is slightly altered by changes in the shape of
the transverse heating profile. The mode coupling strength is strongly affected by
population saturation [10, 12].
In [7] we demonstrate that the threshold of instability can be strongly affected
by amplitude modulation of the pump or signal inputs. The gain of higher order
modes LP11 and LP02 are shown in Fig. 1 of [7]. We show the magnitude of gain
is strongly dependent on the frequency offset and has a dispersion-like curve, and
the peak gain is much larger for LP01 coupling into LP11 than for LP01 coupling
into LP02 . This accounts for the universal observation of transfer into LP11 rather
than into LP02 in experiments.
In [9] we show the relationship the threshold has with amplitude modulation of
the input signal and pump has logarithmic dependence upon modulation depth.
We also show how thresholds are affected by photodarkening and by mode specific
loss (or mode discrimination). For purposes of illustration we treat photodarkening
using a simple model of uniform linear absorption in doped region of the fiber.
In [11], we attempt to match the experimental results presented by Otto et al.
in [3]. We show that the presence of harmonics in the mode content at the output
is indicative of amplitude modulation of the input pump or signal. Power switch-
ing between modes LP01 and LP11 on the millisecond time scale is shown to occur
in the presence of amplitude modulated input signal. The model using the pa-
rameters described in [11] does not produce full power switching between modes,
but we view this as inconsequential since the parameters of the fiber are not fully
described in [3] (e.g. mode discrimination, V parameter, and photodarkening),
and the spectra of pump and input signal noise are likewise unspecified.
In [13] we develop the theory of spontaneous thermal Rayleigh scattering and
show it is stronger than quantum noise so it should be used to seed the modes in
most cases in place of quantum noise.

4. DTU Model
This model was first described in a short paper [14], and later expanded in a longer
paper [15]. Later a study of thermal lensing and its impact on thresholds was added [27].
Assumptions: Arbitrary but fixed mode profiles, frequency-space treatment, broad-
band or narrow band signal, heating assumed to follow the signal irradiance profile,
no population saturation / transverse spatial hole burning, mode coupling gain is
related directly to laser gain rather than pump power, no longitudinal heat flow,
cylindrical thermal boundary, thermal lensing assumes a uniform deposition of the
quantum defect heat across the full core to find a temperature profile to include
when computing thermally lensed modes that are constant in z
Methods: Broad band Greens function temperature solver, coupled mode propagation
method.
Features: Uses coupled mode equations for two modes, LP01 and LP11 ; the mode cou-
pling gain is related to the laser gain; the mode instability threshold is found to
be determined by the laser gain, without direct dependence on other amplifier pa-
rameters; in [14] the threshold for a 1030 nm signal is around 400 W, independent
of pump direction, core and cladding size, etc. (there is a weak dependence on the
V parameter); no run time is reported.
Notes: The frequency dependence of mode coupling gain is studied for different fiber
core sizes. The mode coupling gain vanishes for zero frequency offset.
Table 1 of [14] lists predicted mode instability thresholds for quantum noise ini-
tiation of LP11 . Threshold signal powers are 440 W for a 20 m diameter core;
458 W for a 40 m diameter core; 479 W for an 80 m diameter core. Each case
used V = 3. There is no variation with fiber length or degree of gain saturation.
Increasing V from 3.0 to 3.5 for the 40 m diameter core reduced the threshold
from 458 W to 401 W. The lesson is that the threshold is always near 400 W unless
V is reduced to near the single mode limit of 2.4 or unless the doped diameter is
made smaller than the core. For example, with a core diameter of 40 m and a
doping diameter of 20 m the threshold more than doubles, from 458W to 1035
W.
In a subsequent publication [15], a more comprehensive description of the model
is presented, and they apply it in two new ways - using an input signal with
a Gaussian relative intensity noise, and a narrow-band of amplitude modulated
signal (similar to [7]). They make the claim that an amplitude modulated signal
is very likely the case in any experiment, and it causes the threshold of modal
instability to be much lower than for quantum noise seeding. It also leads to
periodic oscillations in output modal power. This is a claim remarkably similar
to [7].
They vary the convection coefficient of the boundary cooling from hq = 10 W /
(m2 K) to hq = 104 W / (m2 K) and find the mode coupling gain coefficient is un-
changed, in apparent contradiction to [4]. However, they consider only symmetric
cooling.
They vary input signal power: 1 W, 10 W and 50 W. Thresholds increase from
448 W to 480 W and 537 W. This is not too surprising since there will be less heat
deposited per unit length for a fiber that has a gain of 20 dB vs. one that has a
gain of 27 dB, and because in this model the mode coupling gain is proportional
to the signal laser gain.
They vary the amplitude modulation of the LP11 seed light and show an ap-
proximately logarithmic reduction in threshold with modulation amplitude. Con-
sequently only modest improvements on threshold can be expected by reducing
amplitude modulation.
They show that for higher V parameters, the frequency of maximum coupling
increases slightly, but, more importantly, the magnitude of the peak of the coupling
gain increases substantially. For example, for V = 3, peak 0.07 W1 ; for V = 5,
peak 0.09 W1 . They also show that LP01 -LP11 coupling is roughly twice as
large as LP01 -LP02 coupling, which they say explains why LP02 isnt the mode
that mixes with the fundamental in laboratory studies. This agrees with the claims
of [7].
They can show modal power switching well above threshold. Below and near
threshold, the spectrum of LP11 at the output end is very narrow and has one
peak for quantum seeding. The spectrum broadens a bit as gain exceeds threshold.
For amplitude modulation above threshold, this spectrum also shows harmonics,
and well above threshold broadens significantly. Thus, the presence of harmonics
in the spectrum of LP11 is indicative of some starting amplitude modulation, as
also claimed in [7].
In ref. [27] the effect of thermal lensing on modal profiles and resulting effect
on mode coupling is considered. The thermally lensed mode calculation assumed
a uniform distribution of heat across the core. The calculation could have been
made self consistent by iterating the mode and temperature calculations to reach
a converged solution under the assumption that the heat profile is that for the
fundamental mode. The fibers considered had a high degree of index structure
and the guidance properties were sensitive to temperature. The modes modified
by thermal lensing were used to compute the mode coupling gain near the output
end of a counter pumped amplifier. In interpreting the conclusions it should be
remembered that thermal lensing confines the light to a smaller region of the core,
accentuating the influence of population saturation which is not included in the
DTU models.
The DTU results, apart from the thermal lensing results, are all in good qual-
itative agreement with the AS-Photonics results under conditions where direct
comparisons are possible.

5. Clemson Model
This model is described in a paper by Liang Dong [16].

Assumptions: Arbitrary but fixed mode profiles, steady-periodic heating, heating pro-
file from irradiance profile, includes longitudinal heat flow, cylindrical thermal
boundary, cylindrical grid.
Methods: Cylindrical coordinates, semi-analytic temperature solver using tempera-
ture modes, coupled mode method.
Features: The temperature solver uses a cylindrical thermal boundary and includes
longitudinal heat flow, thresholds and gains are compared with the DTU model,
no run time is reported but it should be very fast.
Notes: Dong considers only single-frequency inputs (though he does discuss spectral
considerations in the latter part of his text). Longitudinal heat flow is included in
this model, but it is shown to contribute negligibly to mode coupling.
The threshold reported here is suspect because the starting power in the parasitic
mode is 2 1024 W which is far below the level of quantum noise, and his starting
power isnt sufficiently justified in the text. The thresholds are near 400 W but
they should be closer to 300 W if the starting power were adjusted to quantum
noise levels of 1016 W. This lower threshold would be consistent with the DTU
results if the 50% higher quantum defect associated with 1060 nm signal of this
model compared with the 1030 nm DTU model were used to adjust the threshold.
Dong shows that if the frequency offset is zero, the mode coupling coefficient
disappears.
Dong attempts to describe behavior above threshold by re-seeding the signal light
so once the signal has transferred into LP11 it can couple back to LP01 . One result
of this scheme is shown in his Fig. 6, which is clearly far above threshold. There,
hes turned the amplifier gain up to 34 dB, so his 19.2 W signal is amplified
to somewhere around 50 kW, and notes that the spatial frequency at which the
power transfer occurs is increasing as you progress along the fiber. Then he claims
this behavior and some external perturbation or an unspecified feedback effect
might account for the chaotic behavior seen in experiment in [3]. This claim is
unconvincing.

6. Air Force Models


There are three versions of the Air Force model. Each is applied to transient cases only,
where the boundary conditions are suddenly changed at time zero. The first model was
described in a 2012 paper [4]. A modified formulation of the model was later developed
by Shadi Naderi et al. at AFRL, and is described in a Photonics West paper [23] and
a full article [25]. These two implementations are very similar, but the Naderi model
is reformulated in cylindrical coordinates so true circular boundary conditions can be
applied. A second modified formulation using cylindrical coordinates was published by
Ward [26] in which a numerical beam propagation method is used in place of the original
coupled mode method.

Assumptions: Arbitrary but fixed mode profiles except in Ward revision where arbi-
trary profiles are allowed; transient heating; heating from population inversion;
longitudinal heat flow considered first version; square grid in first version, cylindri-
cal grid in Naderi revision [25] and Ward revision [26]; in the original paper [4] the
pump feedback effect is turned off by artificially fixing the pump longitudinal pro-
file to the initial steady state profile in the initial version; the fiber is co-pumped
in the Naderi version and counter-pumped in the original and Ward revision; the
slinky effect is strong because of the strong transient at time zero in all versions.
Methods: The original version develops the coupled mode method, with coupled mode
equations that include terms for temperature-induced mode coupling and also for
changes in the propagation constant for the individual modes that are computed
from the temperature profile within the core. The temperature solver is unspecif-
ically described as a Crank-Nicolson scheme in the first version; in the Ward
revision the temperature solver is described as a hybrid of finite element and fi-
nite differences; the temperature is solved using an ADI method in the Naderi
revision. The Ward revision develops a new beam propagation method to replace
the coupled mode treatment.
Features: The original version uses PCF modes and a counter-pumped configuration.
That model begins with steady state solution for one set of input pump and signal
powers. At t=0 the input conditions are suddenly changed, with 2.5 W of the total
input signal power of 50 W suddenly transferred from the LP01 mode to the LP11
mode. The Naderi revision starts with a cold, (step index?) fiber and turns on
the full pump power at t=0, with 75-175 W of seed in LP01 plus 5% of that as
unshifted seed light in LP11 . The Ward revision treats a step index fiber, and at
time zero the pump power is ramped from zero to the set point power over 10 ms.
There are no frequency offsets of the input signal in the original version. In
Naderis revision the input signal is sometimes amplitude modulated, produc-
ing frequency side bands in the 500 Hz to 2 kHz range. Run time is several days
in the Naderi revision because the full transverse profile is used in the thermal
solver, and because the model must run long enough for the initial transient to
decay away. Run time of the Ward revision is unspecified, but it is designed to
run on a high power, multi processor computer.
Interpretation (original version): The original model produces mode coupling
without introducing a frequency offset for light in LP11 . However, it is difficult
to discern a distinct threshold from the presented data. The authors do not offer
an explanation of the physical mechanism responsible for initiating or sustain-
ing this mode coupling. Clearly, at early times, before thermal equilibrium is
reached, there are moving gratings as expected from the slinky effect. However,
after many milli seconds equilibrium should be approached, the gratings should
cease to move, and power transfer to LP11 should cease. This behavior appears
to be present in the Naderi revision when operated below threshold (see movie1).
Above threshold mode coupling does not cease after the settling time. Does this
mean there is a physical process that causes mode coupling, or is a frequency shift
somehow introduced by the numerics of the model? How can sustained oscillations
arise in response to only an initial perturbation? Perhaps the authors believe the
conceptual model of Jena [17] can explain this?
We described above in section 3 how numerical truncation errors can form a broad
band noise background similar to quantum noise, and the noise within the gain
band could be amplified to threshold in the Air Force models. There are multiple
tests of the numerical model that can detect this effect. The first is a comparison
of the threshold computed by this model with the thresholds computed by the
AS-Photonics or the DTU or the Clemson models. Is the gain computed using
those models sufficient to amplify truncation noise to threshold when the threshold
power levels of the Air Force model are used? Another test is Fourier transforming
the field amplitudes to search for frequency offsets of the amplified LP11 light at
various points along the fiber.
This paper claims that cooling conditions at the thermal boundary change the
threshold, but in fact there seems to be no such effect on the mode coupling
equations if the overall temperature of the fiber is changed - only the difference
in propagation constants should impact the transient slinky effect. It is true the
mode profiles will change with heating when only one side is cooled, but this
effect is not included in the model since the modes are not allowed to change to
accommodate an asymmetric temperature profile.
Interpretation (Naderi revision): Unlike the original version, the Nadari version
seems to produce well defined thresholds, at least in some cases. Naderi et al.
claim once again that the whole fiber, out to the cooled surface, should be in-
cluded in the thermal model in order to correctly account for thermal boundary
conditions. While this is technically true, it makes the run times exceedingly long,
and the paper never demonstrates that including the full fiber makes a significant
difference in the threshold in any specific cases. The thermal diffusion time for
the 250 m radius of the cooled boundary is approximately 70 ms but the model
appears to produce stable outputs after only 20 ms suggesting the distant thermal
boundary is not involved in an important way. The impact of the outer boundary
location was not tested by varying its diameter.
The Naderi et al. paper contains a number of interesting peculiarities. For one, it is
claimed that no frequency offsets need be involved in creating a mode instability.
However, when the input seed is amplitude modulated, the threshold is reduced
by a factor of two or so. It is obvious that such seed modulation produces the
required frequency shifted side band that is amplified to threshold. The threshold
without this modulation is almost certainly due to amplification of frequency offset
numerical truncation noise, as explained above.
Another peculiarity is the claim that quantum noise level seeding does not trig-
ger mode instability. However, they test this by seeding only the zero frequency
(unshifted) higher order mode with varying powers and, not surprisingly, see no
change in pump threshold power. It is not explained how one can turn off quantum
noise at shifted frequencies in the laboratory, so this treatment violates fundamen-
tal physical laws. If the shifted frequencies were seeded at the quantum level, we
expect the threshold would fall by a factor of approximately two, just as it does
for the case of amplitude modulated seed light.
This paper also contains the claim that steady-periodic models cannot incorpo-
rate multiple frequencies. This is wrong. It also claims that the steady-periodic
models must find the frequency of maximum gain by trial and error, when it is
actually quite easy to predict the frequency [10] within 50 Hz. In any case, the
modulation frequency was varied in this paper to locate, by trial and error, the
frequency offset of maximum gain. The field amplitudes (rather than irradiances)
are not Fourier transformed to look at the spectra of the modes. Unfortunately
this paper, like too many in the field, does not present the full list of amplifier
parameters necessary to perform model comparisons or to replicate the model.
Missing information includes: pump wavelength, signal wavelength, emission and
absorption cross sections for pump and signal, numerical aperture of the fiber,
upper state lifetime, thermal conductivity, heat capacity, mass density, time,
and lengths of the confined doping fibers.
The model is used to compute thresholds for different distributions of Yb3+
dopant. One variant is a simple confinement of the ions to the center of the core.
Another is a three-fold symmetric doping profile. Unfortunately the orientation of
this profile relative to the mode symmetry is not given. Both of the special doping
profiles are found to raise the instability threshold.
Interpretation (Ward revision): The model update by Ward [26] replaces the cou-
pled mode treatment with a beam propagation method based on a one dimen-
sional finite element treatment combined with an angular expansion method. A
new thermal solver is also developed and applied. Cylindrical coordinates are used.
The model is still applied to a transient situation with integrations over the range
0-100 ms. The fiber has a step index profile and step doping profile. The fiber is
again counter pumped only. Longitudinal heat flow is allowed, and the tempera-
ture is computed over the full fiber diameter of 400 m, but the integration time
of 100 ms not sufficient to reach full thermal equilibration with the outer fiber
boundary.
There are no Fourier transforms of the field amplitudes and thus no unambiguous
information about spectral shifts. Frequency shifts of the input signal are included.
Ward sees the threshold power is a factor of three lower than in the original coupled
mode version and attributes this to the inclusion of all possible transverse modes
in place of just two. This seems an unlikely explanation to us but we cannot give a
better one. However, there are no really clear thresholds identified in either case.
Ward recognizes that quantum noise initiation might be important, but doesnt
add noise to the model to test the idea.
This paper strongly restates the original claim that instability occurs only near the
output end of the amplifier. This is probably just the masking effect since the seed
includes strong unshifted LP11 power. He also says the sudden onset of instability
late in the fiber might explain the lower thresholds for larger core diameters,
although he does not offer any support for this claim or for this threshold trend.
The Ward revision still presents plots of the Fourier-transformed irradiance, rather
than Fourier-transformed fields, discarding valuable spectral information. Addi-
tionally, the Fourier transformed data, as presented, may includes the strong ini-
tial transient response to the perturbation, perhaps hiding what may be more
important near-steady-state responses.
Ward offers a good plan for future studies at the end of section 6 of this paper,
and we look forward to these results.

7. Jena Model
This model is described in a 2012 paper [17].
Assumptions: Transient heating; heating from population inversion; cylindrical
boundary; cylindrical grid; cylindrically symmetric fields LP01 and LP02 .
Methods: Cylindrically symmetric BPM; temperature solver unspecified.
Features: This is actually more a conceptual model than an actual numerical model.
They do develop a numerical model of a train of 1 ns pulses, but they run the
model for only 1 ms [17] which is shorter than the time for heat diffusion across
the core. However, they endorse the Air Force numerical model [24] as the most
... so evidently they believe the Air Force model is a good implementation of
their ideas.
The conceptual model is based on zero frequency offset between input signal modes
at the fiber input. Phase shifts are said to develop between the irradiance grating
and the temperature grating due to the rapid temperature rise with z in a counter-
pumped configuration. Longitudinal heat flow is said to enhance this but is not
necessary. This grating shift, or phase shift, is sufficient to cause power transfer
between modes as the irradiance and temperature gratings chase one another
downstream. The resulting Doppler shift is said to cause frequency offsets. Then
the feedback effect accentuates the power transfer and perhaps leads to oscillatory
behavior.
Notes: It is unclear how the proposed mechanism of mode coupling translates to co-
pumped fibers where the temperature gradients are far different from counter-
pumped fibers, yet reportedly have similar instability thresholds. No convincing
evidence is given for the claimed phase shifts, nor is any citation of such an effect
given. The Air Force model has the feedback effect artificially turned off (the
absorbed pump power distribution in z is forced to remain unchanged regardless of
the relative strengths of LP01 and LP11 ), so that model does not fully incorporate
these ideas. Jena seems to hedge on the claim of spontaneous breaking of the
steady state by saying perturbations of some sort may be necessary to initiate
the mode coupling process (no further information is provided about the nature
these proposed perturbations or their source), but they cannot decide whether
their conceptual model would settle into a steady state or be chaotic. A periodic
perturbation is not considered. Cooling conditions should make a difference in this
model because poorer cooling leads to larger longitudinal temperature gradients,
and it is the large temperature gradient that is claimed to start the process.

8. Michigan Model
This model is described in a 2013 paper [18].

Assumptions: Time dependence need not be steady-periodic. The steady-periodic


case is labeled the static case in ref. [18]. The general case is labeled time-
dependent. Frequency offset between LP01 and LP11 is included, unlike in the
Air Force model. The heat profile given by Eq. (5) includes a population inversion
but it does not have transverse profile. The transverse heat profile appears to be
that of the irradiance. Longitudinal heat flow is not included. Single frequency
LP01 and LP11 . Fiber is counter-pumped.
Methods: Coupled mode method, thermal equation solved by separation of variables in
cylindrical coordinates, integration of coupled mode equations is semi-analytical,
pump is turned on at time zero it appears.
Features: When operating slightly above threshold the time-dependent model pro-
duces strong temporal oscillation in the higher order mode even though it is seeded
at a single frequency. In other words, another frequency appears in LP11 and it is
detuned to the blue side of the LP01 frequency. As pump power increases to ten
times threshold more equally spaced frequency sidebands appear and the tempo-
ral oscillation become complicated. A red cascade occurs and shifts the weight of
the LP11 spectrum to several times the maximum gain point. When the detuning
of the seed is not near the point of highest gain the spectra of LP11 become more
continuous and lose most of the discrete sideband structure.
Notes: The red cascade and broadening noted in this paper are qualitatively similar
to the behavior noted by Hansen [15] for multifrequency seed, and by Smith [11]
for amplitude modulated seed or pump, and to a lessor extent for unmodulated
inputs. The threshold is lower because the seed power is 105 W rather than the
1016 W used by Hansen and by Smith. One question is how the harmonic (and
non harmonic) offset frequencies arise in this model? No physical explanation is
given. At low powers (150 W and 500 W) the oscillations are periodic after the
initial transients, so the steady-periodic models should give the same results. Do
they? Another question regards the very high maximum pump level of ten times
threshold. At that point numerical truncation errors can be amplified to threshold.
Is that important here?

9. Comments
Our first comment on alternative models of mode instability is that the AS-Photonics,
DTU, and Clemson models are not wrong. They all start with frequency shifted light
in LP11 and find it grows exponentially to threshold. There can be little doubt that
this gain process is real. Furthermore the three models listed produce thresholds that
are in close agreement when LP11 is seeded with frequency shifted light at the quantum
noise level and when the same assumptions about the heat profile are used. We think
this implies that other models that predict substantially different thresholds are either
incorrect or are not correctly applied.

9.1. Numerical noise


Some researchers maintain that it is not necessary to introduce frequency shifted seed
light in LP11 , even at the quantum noise level. They claim that LP11 grows by some
other physical process with a sharp threshold [4, 17, 25,26]. To test whether this occurs
in our model as well, we ran many cases without frequency shifted seed light. The time
span was set to make the lowest frequency offset 150 Hz, and 30 harmonics were in-
cluded so light in any fiber mode could include a set of frequencies spanning the range
-4.5 kHz to +4.5 kHz, spaced by 150 Hz. We set the pump power to a level well above
the threshold computed when using quantum noise seeding. We found that after the
first propagation step broad band light appeared in LP11 . This low power light which
arose from numerical truncation error which subsequently experienced exponential gain.
After propagating several millimeters the light had a spectrum that closely mimicked
the STRS gain spectrum described in [7, 15]. Apparently truncation errors in the field
variable or temperature provided a low level, broadband seed that was amplified, even-
tually reaching threshold. This pseudo threshold was higher than the quantum noise
threshold by a factor that depends on the ratio of the noise power to the quantum noise
power. This numerical noise also affects the transverse profile because it populates all
of the fiber modes at a low level.
When the power of the LP11 seed light is set to zero, truncation noise adds a seed
power of roughly 1029 W in our model (our model uses double precision variables).
This numerical noise behaves much like quantum noise except the threshold is raised by
approximately 75%, as is to be expected from the lower initial noise level. Because the
truncation noise is broad band, it can seed LP11 in the STRS gain band, and we find
that the frequency spectrum of the output LP11 light mimics the STRS gain profile.
Figure 2 illustrates the amplification of numerical noise to threshold.
Under certain choices for the time span and frequency our model can avoid the
contributions of numerical noise. Then when we seed LP11 with only unshifted light at
powers from zero to several watts, we find there is no modal instability even at powers
of 5 kW which is 10 times the threshold power when quantum noise seeding is included
for this example. The same results are found for co- and counter-pumped amplifiers.
Our model does include all the physical effects that are claimed by the Jena group to
cause mode instability except for longitudinal heat flow [17]. Any diabatic waveguide
effects should express themselves in our model which is based on beam propagation in
the presence of thermally induced index changes. Yet we do not see any such effects
even at 10 times our usual threshold.
It is impossible to state with certainly the numerical noise levels in models other than
our own. However, it is possible that such noise is important if deliberate frequency offset
seeding of LP11 is omitted. The level of truncation in double precision calculations is
1016 . If we assume the field amplitudes initially include this level of noise, the power
noise at the offset frequency might be 1032 W, implying that a power gain of 320 dB
would be required to amplify the noise to threshold. This is twice the gain of 160 dB
required to amplify quantum noise to threshold.
Z = 0 [m]
0
10

10
10
Power [W]

20
10

30
10
6 4 2 0 2 4 6
Frequency [kHz]

Z = 1.6 [m]
0
10
Power [W]

10
10

20
10

6 4 2 0 2 4 6
Frequency [kHz]

Fig. 2. Initial LP01 seed spectrum (upper plot blue trace) and LP11 noise spectrum
(upper plot green trace) when LP11 is seeded by numerical noise alone. The noise
level in LP11 is 1029 W in LP11 . Output spectra at threshold are shown in the
lower plot. The red shifted noise in LP11 has been amplified to threshold. The
LP11 spectrum matches quite well the gain curve for STRS for this fiber.
9.2. Essential numerical tests
Certain modeling tests should be performed on any model to detect the contributions
of numerical error or to identify alternative physical processes. The first test is to add
a small amount of LP11 light frequency shifted for maximum STRS gain. If light at the
quantum noise level lowers the threshold, the higher threshold without such seed light
will not be observable in the laboratory and probably does not have a physical origin.
The second test is to Fourier transform the light from time to frequency at z locations
along the fiber to see whether a frequency offset has developed even in the absence of
deliberate seeding with shifted light. A frequency shift is expected early in the fiber
if numerical noise is important, but not if the proposed alternative physical process is
responsible for mode instability. Another way to detect the frequency shift is to view
movies of the irradiance grating or the temperature grating. The Ward paper present
plots of the Fourier-transformed output power, rather than Fourier-transformed fields,
but such power spectra discard revealing information. Additionally, the transformed
data presented in that paper apparently includes the early transients, perhaps masking
a steady state response.
It is also important to compare thresholds among the different models using the
same input parameters. We have compared the model of Hansen et al. and the model of
Dong with our model and obtain nearly identical results for all three. Unfortunately, we
cannot compare with the Air Force model because many of their inputs are unknown.
Another interesting test of the Air Force model is to start with different initial tran-
sients and see whether similar results are obtained after the transient has decayed to
insignificance. In fact, [2] finds in a laboratory measurement that transient mode cou-
pling dies away after a short time (on the order several seconds) before settling into a
condition with a high fraction of the signal power in the higher order mode. The settling
time should be comparable to the thermal diffusion time to the outer fiber boundary.

9.3. Steady-periodic limitations


Some authors claim our model is limited in some respects by our assumption of steady-
periodic behavior. However, as illustrated in the example shown in Fig. 2, we can set
the time span to a time much larger than the inverse of the maximum gain frequency.
We can also include any number of harmonics of the inverse time span. This allows us
to approximate broad bandwidths with a dense set of frequencies. We cannot model
strong transients such as turning on the pump power at time zero, but observations
indicate that mode instability is a steady state process. There are also transient effects,
discussed under the slinky effect above, but those are not the primary concern of most
users of high power fiber amplifiers.

9.4. Cooling conditions


An important claim of the original Air Force paper [4] is that the cooling conditions
at the fibers thermal boundary can change the mode instability threshold power by a
factor of two or more. It is expected that the transient slinky effect will be stronger
with poorer fiber cooling. However, after the initial transient there should be no effect
on the mode coupling equations if the overall temperature of the fiber is changed. It
is true the mode profiles will change with heating power when only one side is cooled
as in ref. [4], but this effect was not included in the original Air Force model since the
modes were not allowed to change to accommodate an asymmetric temperature profile.
It is not clear whether the Naderi version of the Air Force model also finds the cooling
to be important.

Vous aimerez peut-être aussi